Modeling the temperaturenitrate relationship in the ... - CCE LTER

3 downloads 27511 Views 6MB Size Report
The best model included oxygen, temperature, and salinity (D2 ¼96.6%) and adequately ... coastal upwelling domain of the California Current, J. Geophys. Res. ...... were collected by Pat Wheeler (College of Oceanic and Atmospheric Scien-.
JOURNAL OF GEOPHYSICAL RESEARCH: OCEANS, VOL. 118, 1–17, doi:10.1002/jgrc.20216, 2013

Modeling the temperature-nitrate relationship in the coastal upwelling domain of the California Current Daniel M. Palacios,1,2 Elliott L. Hazen,1,2 Isaac D. Schroeder,1,2 and Steven J. Bograd2 Received 9 January 2013; revised 22 April 2013; accepted 23 April 2013.

[1] Given the importance of nitrate in sustaining high primary production and fishery yields in eastern boundary current ecosystems, it is desirable to know the amounts of this nutrient reaching the euphotic zone through the upwelling process. Because such measurements are not routinely available, we developed predictive models of water-column (0–200 m) nitrate based on temperature for a region of the California Current System (30–47 N) within 50 km from the coast. Prediction was done using generalized additive models based on a compilation of 37,607 observations collected over the period 1959–2004 and validated with a separate set of 6430 observations for the period 2005–2011. A temperature-only model had relatively high explanatory power (explained deviance, D2 ¼ 71.6%) but contained important depth, latitudinal, and seasonal biases. A model incorporating salinity in addition to temperature (D2 ¼ 91.2%) corrected for the latitudinal and depth biases but not the seasonal bias. The best model included oxygen, temperature, and salinity (D2 ¼ 96.6%) and adequately predicted nitrate temporal behavior at two widely separated locations (44 39.10 N and 32 54.60 N) with slight or no bias [root-mean-square error (RMSE) ¼ 2.39 and 0.40 mM, respectively). For situations when only temperature is available, a model including depth, month, and latitude as proxy covariates corrects some of the biases, but it had lower predictive skill (RMSE ¼ 2.50 and 5.22 lM, respectively). The results of this study have applications for the proxy derivation of nitrate availability for primary producers (phytoplankton, macroalgae) in upwelling regions and for biogeochemical and ecosystem modeling studies. Citation: Palacios, D. M., E. L. Hazen, I. D. Schroeder, and S. J. Bograd (2013), Modeling the temperature-nitrate relationship in the coastal upwelling domain of the California Current, J. Geophys. Res. Oceans, 118, doi:10.1002/jgrc.20216.

1.

coastline) was devised in the 1970s as a large-scale estimate of the amount of water upwelled from the base of the Ekman layer [Bakun, 1973; Schwing et al., 1996]. However, UI is derived exclusively from geostrophic estimates of wind stress over the ocean and does not contain information on the properties of the water being upwelled. In order to obtain a more direct indication of the potential for upwelled waters to sustain high biological production, it is desirable to know the amounts of the major nutrients such as nitrate reaching the euphotic zone through the upwelling process. [3] Routine, direct measurements of nitrate, however, are scarce for large regions of the world oceans, and therefore statistical prediction is by necessity undertaken by exploiting the relationship between nitrate and other more widely available measurements such as temperature [e.g., Kamykowski and Zentara, 1986; Garside and Garside, 1995; Louanchi and Najjar, 2000]. The advent of satellite remote sensing in the 1980s and 1990s enabled the production of large-scale maps of surface nutrient concentrations estimated from satellite-derived sea-surface temperature measurements [Traganza et al., 1983; Dugdale et al., 1989, 1997; Sathyendranath et al., 1991; Morin et al., 1993; Goes et al., 1999; Kamykowski et al., 2002; Henson et al., 2003], an area of research still active today [Silio-Calzada et al., 2008; Steinhoff et al., 2010; Sarangi, 2011].

Introduction

[2] Injection of nutrients into the upper layer of the ocean through the process of wind-induced coastal upwelling results in elevated primary production in eastern boundary currents. Although direct measurement of the upwelling process remains elusive, the coastal upwelling index (UI; defined as the magnitude of the offshore component of the Ekman transport, in meter cube per second per 100 m of

Additional supporting information may be found in the online version of this article. 1 Cooperative Institute for Marine Ecosystems and Climate, Institute of Marine Sciences, Division of Physical and Biological Sciences, University of California, Santa Cruz, California, USA. 2 Environmental Research Division, Southwest Fisheries Science Center, National Marine Fisheries Service, National Oceanic and Atmospheric Administration, Pacific Grove, California, USA. Corresponding author: D. M. Palacios, Environmental Research Division, Southwest Fisheries Science Center, National Marine Fisheries Service, National Oceanic and Atmospheric Administration, 1352 Lighthouse Avenue, Pacific Grove, CA 93940-2097, USA. (Daniel.Palacios@ noaa.gov)

©2013. American Geophysical Union. All Rights Reserved. 2169-9275/13/10.1002/jgrc.20216

1

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

[4] A typical temperature-nitrate (T-N) scatterplot shows a strong inverse relationship, such that even a simple linear fit can be statistically significant and have a high coefficient of determination (R2). Upon closer inspection, however, the shape of the relationship may not necessarily be linear, and it may display a broad scatter. A widely used approach for nitrate prediction has been to fit polynomial expansions of temperature, typically quadratic [Chavez et al., 1996; Louanchi and Najjar, 2000] or cubic [Kamykowski and Zentara, 1986; Switzer et al., 2003], to account for potential curvature in the relationship within a linear regression framework. Curve fitting with a sigmoid function has also been used as a form of nonlinear regression [Sarangi, 2011]. The inclusion of covariates such as salinity, chlorophyll concentration, or wind speed in the models tends to further increase the amount of nitrate variance explained [e.g., Roy, 1991; Garside and Garside, 1995; Goes et al., 2000; Sarangi, 2011]. [5] In contrast to least-squares based methods, more recently developed approaches based on local fitting of nonparametric smooth functions of the predictors, such as generalized additive models (GAMs) [Hastie and Tibshirani, 1990; Wood, 2006], provide responses that more closely follow the shape of nonlinear relationships, and therefore have the potential to capture more complex aspects of the T-N relationship. While GAMs do not provide the familiar empirical algorithms estimated by parametric models (i.e., a regression equation), for many applications this is not a requirement since GAMs can be reconstructed from the smoothing function and basis dimension used to fit the model. [6] Our focus here is on the prediction of water-column (0–200 m) nitrate in the California Current System (CCS) within the 50 km region off the coast that is most directly affected by the upwelling process. For this purpose, we build a series of GAMs using a historical data set spanning the period 1959–2004. The resulting models are assessed though residual diagnostics, and their predictive power are evaluated relative to a set of independent observations for the period 2005–2011. As an application, nitrate time series are predicted and evaluated at two sites in the CCS to examine the ability of the models to represent their temporal behavior. To conclude, the strengths and limitations of each model are discussed.

(CoOP/WEST; 1 June 2000 to 20 January 2003; http:// ccs.ucsd.edu/coop/west/), and (4) NOAA’s Newport Hydrographic Line program (NH-Line; May 1996 to present; http://www.nwfsc.noaa.gov/oceanconditions). [8] In addition, we queried historical records from other programs in the World Ocean Database 2009 (WOD09; available at http://www.nodc.noaa.gov/) [Boyer et al., 2009] for this same region, yielding observations for the period 8 July 1959 to 25 July 1985. The WOD09 extract contained data collected in the Monterey Bay area by Hopkins Marine Station (HMS) of Stanford University (14 October 1969 to 22 April 1974) and by Moss Landing Marine Laboratories (MLML) of California State University under the auspices of California Sea Grant (two time periods: 13 April 1970 to 17 December 1971 and 3 January 1975 to 7 December 1976). It also contained observations from the Coastal Transition Zone experiment (CTZ; 30 April to 19 May 1987) off northern California. Observations in the WOD09 for which the accompanying metadata was insufficient to determine the specific program under which they were collected were grouped together under ‘‘other.’’ CalCOFI data contained in the WOD09 were excluded to avoid duplication. [9] Because the focus of our effort was on the waters directly influenced by coastal upwelling, we limited the extracts to observations in the upper 0–200 m in the water column occurring within a strip 50 km from the coastline. For this purpose the great-circle distance between each observation and the nearest coastline was computed using the high-resolution level of the Global Self-consistent, Hierarchical, High-resolution Shoreline Database (available at http://www.ngdc.noaa.gov/mgg/shorelines/gshhs.html) [Wessel and Smith, 1996]. For each source, we retained observations that contained at least three variables: temperature, salinity, and nitrate (in addition to date, depth, and geographic location). If other variables, such as oxygen, phosphate, or silicate were available, these were also retained. The nitrate-to-phosphate (N:P) and the silicateto-nitrate (Si:N) ratios were computed for observations containing these measurements. These variables were used at the different stages of data screening as filtering criteria. Tabular and graphical summaries of the temporal and spatial coverage of the observations used in this study are given in Table 1 and Figure 1.

2.

3.

Data Sources

Data Preparation

[10] Considering the wide variation in data sources, time periods, and sampling protocols in the initial data compilation, quality control was conducted to identify and deal with potentially problematic observations, as is commonly performed on hydrographic databases prior to analysis [e.g., Louanchi and Najjar, 2000]. We implemented the following steps for data screening, outlier identification, and potential transformation.

[7] The CCS is one of the world’s most intensely sampled ocean ecosystems. Multiple field programs and process studies have been conducted in the CCS by both academic and government institutions [e.g., Bograd et al., 2003; Huyer et al. 2007; Pe~na and Bograd, 2007; Checkley and Barth, 2009]. We obtained hydrographic data for the CCS covering the region 30–47 N, 126–116 W from four of these programs: (1) the California Cooperative Oceanic Fisheries Investigations program (CalCOFI ; 7 January 1969 to 29 January 2011; available at http://www. calcofi.org/), (2) the U.S. GLOBEC Northeast Pacific Long-Term Observation Program (LTOP ; 19 September 1997 to 1 September 2004; available at http://globec.whoi. edu/jg/dir/globec/nep/ccs/ltop/), (3) the Coastal Ocean Processes/Wind Events and Shelf Transport program

3.1. Data Screening [11] The initial data compilation contained 54,487 observations in the upper 200 m and within 50 km from the coast. It consisted of 12 columns for date, latitude, longitude, observation depth, temperature, nitrate, salinity, oxygen, phosphate, silicate, N:P ratio, and Si:N ratio (the 2

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Table 1. Sources of Hydrographic Data, Number of Stations, Number of Observations, and Dates in the Cleaned Data Set (See Section 3 for Details on Data Preparation) Programa

Number of Stations

Number of Observations

Start Date

End Date

CalCOFI HMS MLML CTZ LTOP WEST NH-Line Other Total

2731 321 432 11 362 637 102 970 5566

27,544 1755 1826 118 3370 4424 102 4898 44,037

7 Jan 1969 14 Oct 1969 13 Apr 1970 30 Apr 1987 15 Nov 1997 1 Jun 2000 16 Nov 1997 8 Jul 1959 8 Jul 1959

29 Jan 2011 22 Apr 1974 7 Dec 1976 18 May 1987 1 Sep 2004 20 Jan 2003 8 Jun 2011 25 Jul 1985 8 Jun 2011

a

See section 2 for details of the programs.

values of nitrate, phosphate, or silicate equal to zero (n ¼ 3814). The rationale for this being that once a nutrient becomes depleted any temperature-nutrient relationship breaks down and for these observations the temperature at

abbreviated names for these variables, as used in the models, are defined in Table 2). The first step in data screening consisted of removing duplicate entries (n ¼ 84). The second step involved removing observations with measured

Figure 1. (top) Time-latitude plot showing the coverage of the hydrographic data over the 53 year period 1959–2011, colored by major program. The number of observations collected by each program is indicated in parentheses. Vertical lines indicate the breaks used in the subperiods in the bottom. (bottom) Maps showing the location of hydrographic stations in the cleaned data set for the 53 year period 1959– 2011, colored by major program. Data are split into three subperiods for presentation. The number of observations in each subperiod is indicated in parentheses. 3

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Table 2. Descriptive Statistics for the Cleaned Variablesa Variable

Abbreviation

n

Range

Mean

sd

Skewness

Kurtosis

CV

Nitrate (mM) Temperature ( C) Salinity (psu) Oxygen (mL/L) Phosphate (mM) Silicate (mM) N:P ratio Si:N ratio Obs. depth (m) Month Latitude ( N)

N T S O P Si N:P Si:N Z M L

44,037 44,037 44,037 35,786 39,378 42,830 39,378 42,830 44,037 44,037 44,037

[0.02, 37.70] [6.09, 20.62] [31.060, 34.365] [0.79, 9.28] [0.01, 3.22] [0.100, 67.410] [0.051, 60.000] [0.015, 60.000] [0, 200] [1, 12] [30.342, 47.000]

14.48 11.14 33.51 4.41 1.30 18.27 8.94 2.73 56.51 5.63 36.63

10.33 2.28 0.48 1.52 0.67 12.89 4.28 4.98 51.44 3.23 4.14

0.10 0.78 1.61 0.07 0.09 0.60 0.78 4.34 1.06 0.15 1.15

1.70 3.52 6.26 1.93 1.84 2.50 2.8 24.30 3.32 1.92 3.05

71.30 20.49 1.42 34.46 51.65 70.55 47.88 182.46 91.02 57.42 11.31

a The variables used in data screening and in GAM model specification, sample size (n), range, mean, standard deviation (sd), skewness, kurtosis, and coefficient of variation (CV). Note that the second column provides the abbreviated names of the variables used in the text.

water mass occupied much of the water column, and it was generally characterized by relatively cool (T < 10 C), salty (S > 32), low-nitrate (N < 6.5 mM), high-phosphate (P > 1 mM), and low-oxygen (O < 6 mL/L) conditions. While a full study of this water mass remains to be conducted, it was judged that these observations should be removed prior to analysis because their characteristics were not representative of the conditions encountered in this area at any other time in the data set. To accomplish this, a linear regression of phosphate on nitrate was fitted to the nonanomalous observations, and the 95% prediction band (simultaneous, protected against multiple observations with Scheffe adjustment) for this regression was applied to the entire data set as the basis for identifying and excluding outliers (n ¼ 1914) in the nitrate-phosphate relationship. 3.2.4. Multivariate Outliers [15] The final step in the outlier analysis took advantage of concurrent measurements in addition to temperature and nitrate available for several of the data sources. Multivariate outliers were identified following the technique outlined in Tabachnik and Fidell [1989, 2001]. Briefly, the Mahalanobis distance was computed for each observation from the centroid of the n-dimensional point cloud, and a 2 test was used to identify points with distances greater than the critical value of the distribution [with  ¼ 0.001 and degrees of freedom (df) equal to the number of variables included in the particular combination of data points being considered]. This procedure was implemented on two subsets of the data, one containing concurrent observations for the eight variables temperature, nitrate, salinity, oxygen, phosphate, silicate, N:P ratio, and Si:N ratio (n ¼ 35,949), and another containing observations for which only the three variables temperature, nitrate, and salinity were available (n ¼ 10,138). [16] To a large extent the multivariate outliers identified in these two subsets (n ¼ 1847 and n ¼ 198, respectively) occurred near the margins of the n-dimensional point cloud as gleaned from inspection of the respective propertyproperty plots. Closer examination revealed that they could be grouped into three types. Type 1 outliers (n ¼ 926) were mostly in the upper 50 m and in the southern half of the latitudinal range (L < 38 N). These waters had a wide range in temperature (8.25 < T < 24.1 C), a high salinity (S > 33.15), were low in nitrate (N < 20 mM), and high in oxygen (O > 5 mL/L). Type 2 outliers (n ¼ 146)

which depletion occurred is unknown (cf., the ‘‘nitrate depletion temperature,’’ defined as the intercept in a T-N scatterplot at the temperature axis) [see Switzer et al., 2003]. The third step involved removing out-of-range values in temperature (5 < T < 25 C), salinity (30 < S < 35), nitrate (0 < N < 50 mM), phosphate (0 < P < 10 mM), silicate (0 < Si < 150 mM), N:P ratio (0 < N:P < 60), and Si:N ratio (0 < Si:N < 60) for typical oceanic waters in the upper 200 m of the CCS (n ¼ 2279).

3.2.

Outlier Analysis

3.2.1. Univariate Outliers [12] Rather than the traditional boxplot, we used Cleveland dotplots as an efficient graphical tool to visualize univariate outliers. In a Cleveland dotplot, the row number of an observation is plotted against the observation value, and points that stick out on either side are potential outliers [Zuur et al. 2009]. We generated Cleveland dotplots for temperature, nitrate, salinity, and oxygen using latitude and depth as sort variables for the rows, and regarded as outliers observations that departed noticeably from the point cloud (likely caused by measurement or data entry error). The combined univariate outliers from these plots (n ¼ 120) were excluded from the data set. 3.2.2. Outliers in the Temperature-Salinity Relationship [13] Unusual relationships in the temperature-salinity scatterplot, evident as ‘‘tendrils’’ emanating from the main point cloud, were explored. In all cases, these observations (n ¼ 194) were isolated to individual profiles and may have been the result of extreme environmental conditions (such as strong El Ni~ no events in 1970–1971, 1981, 1997–1998) or instrumental error. These observations were excluded from the analyses. 3.2.3. Outliers in the Nitrate-Phosphate Relationship [14] A nitrate-phosphate scatterplot for those observations containing both measurements (n ¼ 43,331) revealed that most of the data set had a well-behaved linear relationship. However, a small cluster with a much steeper relationship (i.e., low nitrate, high phosphate conditions) was readily evident. Further investigation indicated that these observations belonged to a unique water mass found off Oregon and Washington during the period 1961–1970. This anomalous 4

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

and Figure 1). Descriptive statistics for this data set, including range, mean, standard deviation, skewness, kurtosis, and coefficient of variation are provided in Table 2. We also examined the univariate distribution of the main variables (nitrate, temperature, salinity, oxygen, depth, month, and latitude) as well as the bivariate relationships among them and these are shown in Figure 2. Pearson correlation coefficients among variables are presented in Table 3. Phosphate and silicate covaried strongly with nitrate (r ¼ 0.98 and 0.95, respectively) but were also highly collinear with each other (r ¼ 0.95) as well as with oxygen (r ¼ 0.92 and 0.87, respectively) and with temperature (r ¼ 0.83 and 0.82, respectively) ; therefore, they were not included as potential explanatory variables in the models. [19] Both the descriptive metrics and the graphical exploration (Table 2 and Figure 2) indicated some degree of nonnormality in the primary variables used in modeling (N, T, S, O); however common data transformations (i.e., log, square root, inverse) did not improve the descriptive metrics or the distributions. (Although normality can be achieved with more drastic transformations like the BoxCox or the rank, difficulties with interpretability and back

corresponded to a small cluster found at middepths (20– 200 m) and mostly in the northern half of the study area (L > 36 N). These waters were cold (T < 8.25 C) and had a high salinity (S > 33.15), high nitrate (N > 20 mM), and low oxygen (O < 4 mL/L). Type 3 outliers (n ¼ 770) were found mostly in the upper 50 m and occurred primarily in the northern half of the study area (L > 36 N). These waters had a wide range of temperature (6 < T < 17 C), a low salinity (S < 33.15), low nitrate (N < 20 mM), and high oxygen (O > 4.5 mL/L). [17] Thus, these outliers identified the most extreme nutrient relationships in multivariate space and were separable by water mass characteristics, latitude, and depth. The procedure also eliminated observations with salinities lower than 31.060, which are influenced by riverine discharge (e.g., the area of influence of the Columbia River plume as well as other rivers discharging along the Oregon and California coasts). For these reasons, we felt justified to remove these outliers from data compilation. 3.3. Final Data Set [18] The cleaned data set contained 44,037 observations and spanned the period 8 July 1959 to 8 June 2011 (Table 1

Figure 2. (a–g) Univariate frequency histograms (n ¼ 22 bins) for variables used in GAM modeling after screening and outlier analysis. (h–x) Binned bivariate scatterplots (hexagonal bins, n ¼ 30) with color shade indicating the density of observations (blue ¼ low, red ¼ high) for selected variable combinations. The range of the number of observations per bin is given above each figure. Axis units are as in Table 2. 5

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Table 3. Correlation Matrix (Pairwise Pearson Correlation Coefficients) of Main Variables in the Compiled Data Set Considered in Data Screening and in GAM Model Specification

N T S O P Si Z M L

N

T

S

O

P

Si

Z

M

L

1.000 0.824 0.620 0.929 0.983 0.951 0.651 0.059 0.089

1.000 0.255 0.659 0.832 0.819 0.483 0.031 0.428

1.000 0.663 0.556 0.529 0.520 0.028 0.528

1.000 0.921 0.868 0.802 0.024 0.157

1.000 0.951 0.673 0.044 0.125

1.000 0.572 0.064 0.258

1.000 0.004 0.226

1.000 0.083

1.000

denitrification generally did not occur in the upper 200 m anywhere in the data set.

transformation made them less applicable for our purpose). Therefore, we used the variables untransformed and note that some nonnormality remained in the final models (see supporting information for additional diagnostics of model residuals). We also point out that after screening and deletion of outliers the final data set reflected waters of oceanic character and therefore the inferences drawn from the models in the following sections are only applicable to the ranges and variable combinations in the data analyzed here.

5.

Modeling Approach

[24] Although global methods (i.e., linear regression, including polynomial expansions) have been widely used to model the T-N relationship [Kamykowski and Zentara, 1986; Garside and Garside, 1995; Chavez et al., 1996; Louanchi and Najjar, 2000], our focus here is on local methods (i.e., GAMs), which have the potential to better capture nonlinearities in the relationship that may be related to regional gradients (e.g., water mass distributions, upwelling intensity). A brief introduction to the GAM methodology is given in section 5.1 followed by details of variable and smoothness selection in section 5.2. Models quantifying the T-N relationship are built in sections 5.3, starting from a T-only model to more complex models that further explore the nonlinear dependence of nitrate on other influential variables and that address some of the shortcomings of the simpler models.

4. Nitrate Patterns in the Coastal Upwelling Domain of the CCS [20] The large observational data set compiled here allowed us to examine the patterns of biologically available nitrate in the water column in the coastal upwelling domain of the CCS over a period of five decades. As occurs throughout the world ocean, nitrate concentrations in the study area were high at depth and became progressively lower in the upper levels of the water column (Figures 2m and 3a). Nitrate increased with latitude, especially at depth (Z > 50 m) (Figure 3a). In waters Z < 50 m nitrate tended to peak at 37–40 N and then decrease. [21] Nitrate had a strong inverse relationship with temperature and oxygen and a direct one with salinity (Figures 2s–2u and 4a). These relationships were reasonably linear and tight at Z > 50 m, but they developed a ‘‘hockey stick’’ shape and displayed wide scatter at shallower depths, especially where N < 5 mM, T > 14.5 C, S < 33.25, and O > 6 mL/L (Figures 3b–3d and 4a). [22] In terms of temporal patterns, nitrate followed a seasonal cycle with a broad peak in the spring and summer months (Figure 4b). This cycle occurred at all depths although it was less marked in very shallow (Z < 10 m) and very deep (Z > 100 m) waters. There was also evidence of decadal trends over the period 1959–2011, with nitrate decreasing from 1959 to the late 1980s and then increasing through the present time (Figure 4c). This trend occurred at all depths, but, like the seasonal pattern, it was most marked at middepths (10 < Z < 100 m) (Figure 4c). [23] Progressively lower nitrate and oxygen concentrations toward the southern half of the study area (L < 37 N; Figures 2h, 2k, and 3a) indicated that denitrified waters are upwelled in this region relative to the waters upstream. However, the fact that the observed nitrate concentrations were always high at low oxygen levels indicated that

5.1. General Form of a GAM [25] GAMs are a nonparametric extension of the more familiar generalized linear models (GLMs). Instead of assuming a priori any rigid parametric form, GAMs represent the relationship between the response and the explanatory variables by smooth functions, which can take virtually any form [Hastie and Tibshirani, 1990; Wood, 2006]. Because they implement a local, data-driven regression, GAMs can be used to quantitatively explore complex relationships when little is known about the underlying mechanisms responsible for generating the observations. Like GLMs, GAMs are estimated using the method of penalized likelihood and the discrepancy between the observations and the estimated mean is measured using deviance residuals (expressed as the percent of deviance explained, D2) [Hastie and Tibshirani, 1990; Wood, 2006]. [26] The general form of a GAM is: g ð i Þ ¼  þ

p X

fj ðXi Þ þ "i

ð1Þ

j¼1

where the function g(m) is a link function relating the mean of the response variable given the explanatory P variables, m ¼ E(YıX1, . . . ,XP), to the additive predictor  þ fj (Xi). 6

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Figure 3. Trellis scatterplots of N against L, T, S, and O conditioned by Z, with observations grouped into five levels having roughly the same number of counts. The depth levels (in meter) are displayed on the strip above each figure, both numerically (bracketed intervals) and visually (darker shaded sections of the strip). Red curve in all figures is a loess scatterplot smoother (degree ¼ 2, span ¼ 3/4) intended to guide the eye through the point cloud. parameters associated with each smooth term, based on generalized cross-validation (GCV). Simply put, cross validation involves leaving one of the data points out, fitting the model to the remaining data, and then calculating the square difference between those points and the fitted model (smaller differences mean better models). This procedure is repeated for all data points and for several amounts of smoothing (and hence several values of df for each term). The GCV score reflects the overall balance between the gains obtained by increasing the amount of smoothing, and the costs in terms of increasing the number of df [Wood,

The term  represents any strictly parametric component in the model (e.g., the intercept), while components fj (Xi) in the additive predictor are specified as nonparametric smooth functions of the explanatory variables, and "i are independent and identically distributed normal random variables [Hastie and Tibshirani, 1990; Wood, 2006]. [27] Model fitting was carried out in the R environment version 2.15.1 [Ihaka and Gentleman, 1996; R Core Team, 2012] using the ‘‘mixed GAM computation vehicle’’ (mgcv) library version 1.7–22 [Wood, 2006]. The mgcv library implements an automatic selection of the smoothing 7

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Figure 4. Trellis scatterplots of (a) N against T conditioned by S, (b) N against M conditioned by Z, and (c) N against date conditioned by Z, with observations grouped into five levels having roughly the same number of counts. Other plot details are as in Figure 3. while data for the period 2005–2011 (n ¼ 6430) were saved for model testing. This allowed us to evaluate the ability of the models to predict new nitrate observations while also allowing us to assess the models’ representation of temporal behavior at selected locations over the testing period (see section 6). As mentioned in section 3.3, nitrate concentration was treated as a continuous normal random variable, and thus GAMs were built using the identity function as the link function (i.e., N  Gaussian). Thin-plate splines were used as the basis for the smooth function for the explanatory variables except for month (M), which was treated as a cyclical variable using the cyclic cubic regression spline. [29] Selection of the smoothing parameter associated with each term involved specifying the basis dimension, k, which amounts to setting the maximum possible df allowed for each term (the actual effective df for each term are then

2006]. The maximum number of df of each smooth term (and hence the complexity of the relationship) must be set initially by the user, but the fitting procedure subsequently ‘‘downgrades’’ them to minimize the GCV score of the entire model [Wood, 2006]. Optimal smoothness for each term can be automatically achieved by initially fitting a model with restricted maximum likelihood (REML) instead of GCV, but the user must still decide if the functional responses obtained with this method are an appropriate representation of the underlying processes, and make adjustments to the df as necessary [Wood, 2006]. 5.2. Model Specification and Selection [28] The data set was divided into training and testing portions using the following scheme: observations for the period 1959–2004 (n ¼ 37,612) were used for model training 8

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Table 4. Assessment of Single-Term GAM Models Used to Determine the Relative Importance of the Explanatory Variables Based on GCV Score, AIC, Drop in Deviance, and the Percent Deviance Explained (D2)a Model N  Intercept L M Z S T O

df

p Value

GCV Score

AIC

Resid. Deviance

Deviance Drop

D2 (%)

1.00 5.61 4.85 5.82 6.00 5.96 5.97

32.5 that then leveled out at higher temperatures and lower salinities. In addition to this large-scale gradient, a local maximum in nitrate was predicted at low temperature and low salinity (Figure 6a), corresponding to estuarine influences at the northern part of the study area. This model largely corrected the depth and latitude biases in the T-only model, indicating that nitrate levels in the CCS are highly dependent on water mass, but the tendency to overpredict in the

Ni ¼  0 þ f1 ðOi Þ þ f2 ðTi  Si Þ þ "i

ð4Þ

[41] Automatic smoothness selection with REML suggested 8.28 effective df for the O smooth term but the functional response appeared slightly overfitted, so for the final fitting with GCV the basis dimension for this term was constrained to k ¼ 8. The basis dimension for the T  S term was maintained as k ¼ 5 for both T and S, as in the TS model in the previous section. This model used a total of 31.1 df (1 for the intercept, 6.27 for the O term, and 23.83 for the T  S interaction), with all smooth terms being highly significant (p values < 0.001). The fit statistics for this model were the best of any model, with D2 ¼ 96.6%, GCV score ¼ 3.29, and AIC ¼ 118366.7 (Table 5). [42] The smooth term for oxygen, the strongest effect, consisted of a sigmoid curve descending from high nitrate at the lowest oxygen concentrations to low nitrate at the higher oxygen concentrations (Figure 7a). This smooth captured the well-known influence of oxygen levels on 11

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Figure 6. (a) Estimated effects (contoured surface) at the scale of the linear predictor for the T  S interaction term for a GAM based on T and S for the training data set (1959–2004, n ¼ 37,607). (b) Model residuals versus Z (colored by M). (c) Model residuals versus L (colored by M). Red curve in Figures 6b and 6c is a loess scatterplot smoother (degree ¼ 2, span ¼ 3/4). (d) Scatterplot of observed versus predicted values (colored by Z) by a GAM based on T and S for the testing data set (2005–2011, n ¼ 6430). Dashed gray line is the linear fit and black line is the 1:1 line. Further diagnostics of the model’s residuals indicated that, while present at low levels, the issues of nonconstant variance, nonlinearity and departure from normality in the simpler models were greatly reduced with the expanded model and were no longer of concern (Figures S3 and S4 in supporting information). [43] The performance of this model at predicting new observations was evaluated using the testing set (Table 6). This model achieved the highest correlation between the observed and the predicted nitrate (r ¼ 0.99). More importantly, the scatterplot showed that the slope of the relationship was indistinguishable from a 1:1 relationship and that

nitrate uptake, respiration, and microbial processes in the water column [e.g., Sarmiento and Gruber, 2009], as evidenced by the depth-dependent pattern in the partial residuals (colored dots in Figure 7a). The functional response of nitrate to the T  S term was similar to that of the TS model, but temperature tended to be a stronger driver of the relationship than salinity in this model, especially at high nitrate values (Figure 7b). This model also predicted the local maximum in nitrate at low temperature and low salinity (Figure 7b) described in the previous section. No latitudinal bias was apparent in the model’s residuals, and the seasonal bias was also largely corrected (Figure 7c). 12

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

Figure 7. (a) Estimated effects (solid black curve) at the scale of the linear predictor for the O smooth term for a GAM based on O, T, and S for the training data set (1959–2004, n ¼ 29,378). The 95% confidence limits (strictly Bayesian credible intervals) are shown as dashed black lines. Dots are the partial residuals, colored by Z. (b) Estimated effects (contoured surface) at the scale of the linear predictor for the T  S interaction. (c) Model residuals versus Z (colored by M). Red curve is a loess scatterplot smoother (degree ¼ 2, span ¼ 3/4). (d) Scatterplot of observed versus predicted values (colored by Z) by a GAM based on O, T, and S for the testing data set (2005–2011, n ¼ 6408). Dashed gray line is the linear fit and black line is the 1:1 line.

oxygen and salinity, indicating that these three variables captured the dominant physical and biological processes driving nitrate variability in the upper 200 m of the CCS. However, despite its shortcomings, a T-only model may be the only option in applications for which no other properties are available. Thus, we explored models that incorporated depth, month, and latitude as proxies for the spatial and temporal processes accounted for by the OTS model.

there was very little bias relative to depth in the predictions (Figure 7d). 5.6. A Proxy Model With Temperature, Depth, Month, and Latitude [44] The prediction biases in depth, latitude, and season detected in the T-only model (as well as the strong structure in the residuals) were adequately corrected by including 13

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

highly significant (p values < 0.001). The fit statistics were similar to those of the TS model (D2 ¼ 87.2%, GCV score¼ 13.69, AIC ¼ 205112.4; see Table 5). [46] The functional response of nitrate in T  Z space showed a rapid decrease at T < 14 C at all depths, indicating that temperature was a stronger driver of the relationship than depth, and it then leveled out at the higher temperatures that occurred at Z < 100 m (Figure 8a). The smooth term for month (Figure 8b) predicted negative nitrate values from December through May and positive values from July through November, indicating the periods when this term corrected for overprediction and underprediction, respectively, relative to the T-only model. However, this term only made a modest contribution to nitrate prediction, as evidenced also by the wide scatter in the partial residuals (Figure 8b). Finally, the response of nitrate to

Using the previously defined training and testing data sets, the best fitting model was: Ni ¼  0 þ f1 ðTi  Zi Þ þ f2 ðMi Þ þ f3 ðLi Þ þ "i

ð5Þ

[45] Automatic smoothness selection with REML suggested 19.76 effective df for the T  Z interaction, 7.76 df for the M term and 8.46 df for the L term. However, the functional responses for both M and L appeared slightly overfitted, so for the final fitting with GCV the basis dimension for these terms was constrained to k ¼ 7 and 8, respectively, while the basis dimensions for the T  Z term were set to k ¼ 6 and 4, respectively. This model used 1 parametric df for the intercept, 19.51 df for the T  Z interaction, 4.95 df for the month and 6.99 df for the latitude smooth term, for a total of 32.45 df. All three smooth terms were

Figure 8. (a) Estimated effects (contoured surface) at the scale of the linear predictor for the T  Z interaction term for a GAM based on T, Z, M, and L for the training data set (1959–2004, n ¼ 37,607). (b and c) Estimated effects (solid black curves) at the scale of the linear predictor for the M and L smooth terms, respectively, for a GAM based on T, Z, M, and L for the training data set (1959–2004, n ¼ 37,607). The 95% confidence limits (strictly Bayesian credible intervals) are shown as dashed black lines. Dots are the partial residuals, colored by Z in Figure 8b or by M in Figure 8c (note that the y axis has been constrained in these two plots to highlight the functional shape of the smooth functions). (d) Model residuals versus Z (colored by M). (e) Model residuals versus L (colored by M). Red curve in Figures 8d and 8e is a loess scatterplot smoother (degree ¼ 2, span ¼ 3/4). (f) Scatterplot of observed versus predicted values (colored by Z) for a GAM based on T, Z, M, and L for the testing data set (2005–2011, n ¼ 6430). Dashed gray line is the linear fit and black line is the 1:1 line. 14

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

nitrate at seasonal and interannual scales. For this purpose, we inspected time series of predicted nitrate at 150 m depth from the testing data set (2005–2011) and assessed their skill relative to the observed series (via the root-meansquare error, RMSE) for two locations, one on the NH-Line and the other in the CalCOFI domain. These locations corresponded to station NH-25 (44 39.10 N, 124 390 W), located 46 km from shore on the NH-Line, and CalCOFI station 93.3.28 (station 28 on line 93.3; 32 54.60 N, 117 23.40 W), located 13 km from shore. [50] Mean nitrate concentration at NH-25 (31.51 mM) was appreciably higher than at CalCOFI station 93.28 (25.80 mM), but otherwise both stations exhibited a similar seasonal cycle, with lowest levels in winter and fall and highest levels in spring and summer. The predicted time series by all three models captured the observed seasonality at both sites (Figures 9a and 9b). They also captured important interannual variations as seen in the high value in 2008 (spring for CalCOFI and summer for NH-25) and the low value in the winter of 2010 (Figures 9a and 9b). The high values of nitrate for both CalCOFI and NH-25 in spring/ summer of 2008 were likely associated with the La Ni~na event of 2007–2008, when cool sea surface temperatures, strong upwelling, and high primary production prevailed in the CCS [McClatchie et al., 2009]. Conversely, the low nitrate values in winter 2010 occurred during the short and weak El Ni~no event of 2009–2010 [Bjorkstedt et al., 2011]. [51] However, only the OTS model had a high predictive skill (RMSE ¼ 2.39 mM and 0.40 mM, respectively for NH25 and CalCOFI station 93.3.28). The T-only model had the lowest predictive skill (RMSE ¼ 4.13 and 6.39 mM, respectively for NH-25 and CalCOFI station 93.3.28). The

latitude was described by a large-scale gradient, with nitrate decreasing with increasing latitude (Figure 8c). This term corrected for overprediction in the T-only model, especially at the higher latitudes. The two local maxima near 33–35 N and 38–40 N along this gradient (Figure 8c) probably indicate localized areas of increased nitrate supply due to intense upwelling centered around Point Conception and Cape Mendocino, respectively. [47] The model’s residuals indicated that the depth and latitudinal bias were largely corrected but there was evidence that some seasonal bias still remained (Figures 8d and 8e), probably because of the wide scatter in the relationship of this variable with nitrate (Figures 4b and 8b). Further examination of the residuals indicated an improvement over the T-only model, with levels of nonnormality and heterogeneity similar to those of the TS model (Figure S5 in supporting information). [48] The performance of this model at predicting new observations was evaluated using the testing set, with most metrics being very similar to those of the TS model (Table 6). A scatterplot of observed versus predicted nitrate had a slope that departed moderately from the 1:1 relationship model and the shape of the relationship had a slight curvature (Figure 8f), leading to underprediction at intermediate nitrate levels and to overprediction at both low and high nitrate levels (Figure 8f).

6.

Application: Predicting Nitrate Time Series

[49] We conducted further evaluation of three of the models developed in section 5 (T-only, OTS, and TZML models) to investigate the temporal behavior of predicted

Figure 9. Time series of observed and predicted N in the testing data set (2005–2011) at 150 m depth for: (a) hydrographic station NH-25 located off Newport, Oregon, and (b) CalCOFI station 93.3.28 located near San Diego, California. For the two locations the observation data (black line) are compared to three predictive GAM models: T-only (blue line); T, Z, M and L (green line) ; and O, T, and S (red line). RMSE values for each of the predicted series are given in the text. 15

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP

in the testing data set was very high (r  0.97–0.99; Table 6). However, examination of the residuals and the skill at predicting individual time series exposed important biases in most models. The OTS model was the only one that yielded residuals that approached normality and that contained no significant structure (see supporting information). Seasonal bias was present in all models to varying degrees, and it was the least tractable source of bias due to the wide scatter in the relationship between nitrate and month. To some extent, these issues arise because of the varying degree of non-normality present in the input variables (Figures 2a–2d). But if the goal is to produce unbiased nitrate predictions, this study highlights the value of thorough assessment of residuals as a tool for model improvement over other metrics of model performance. [56] Accurate estimation of nitrate concentration in the euphotic zone from proxy variables has several applications. For example, a nitrate inventory for the water column combined with information about thermocline depth and water-column stratification [Palacios et al., 2004] could be used to derive a more direct index of the biological utility of upwelled waters relative to existing wind-derived upwelling indices like the UI. This is not only relevant for the monitoring of primary production mediated by phytoplankton [Saba et al., 2011] and benthic macroalgae [Broitman and Kinlan, 2006] but also for the estimation of secondary production and fisheries yields [Friedland et al., 2012]. Biogeochemical and ecosystem numerical models that use nitrogen as currency [e.g., Powell et al., 2006; Doney et al., 2009; Somes et al., 2010] could also benefit from the statistical relationship between nitrate and variables like temperature, salinity, depth, and latitude, to constrain their nitrate budgets and thus obtain improved estimates of phytoplankton primary production and biomass. [57] Finally, large-scale prediction of surface nitrate has relied on satellite measurements of sea surface temperature and chlorophyll concentration [e.g., Goes et al., 1999; Kamykowski et al., 2002, Silio-Calzada et al., 2008, Sarangi, 2011]. Given the improvement provided by our GAM models that included salinity over the T-only model, it is expected that the incorporation of satellite-measured sea surface salinity, which only recently became available [Lagerloef, 2012], into these efforts will result in improved maps of surface nitrate at regional and global scales.

proxy TZML model significantly improved the skill for the CalCOFI station (RMSE ¼ 2.50 mM) but not for NH-25 (RMSE ¼ 5.22 mM). [52] In spite of the corrective steps taken to minimize the biases in the T-only model with the TZML proxy model, it still underpredicted nitrate at both stations (Figures 9a and 9b). This is not surprising, considering the wide scatter of the residuals relative to the fitted responses for both month and latitude (Figures 8b and 8c). In contrast, the OTS model showed good agreement with the observations, although the slight underprediction at NH-25 (2.39 mM RMSE) suggested that a regional bias remained, most likely due to the majority of the training data coming from southern California (see Table 1 and Figures 1 and 2), where nitrate levels tended to be less variable (Figure 3a).

7.

Discussion and Conclusions

[53] The flexible GAM framework provided optimal nonlinear functional responses describing large-scale vertical and latitudinal gradients in nitrate in the CCS, as well as more complex interactions related to water mass distributions that could not be easily replicated with parametric models. However, the usefulness of GAMs to predict nitrate at larger spatial scales (basinwide to global) has not been evaluated, and future studies should compare their performance relative to traditional regression approaches, which have relied on implementing separate subregional models in order to improve predictive power [e.g., Chavez et al., 1996; Goes et al., 2000; Louanchi and Najjar, 2000]. [54] The four models considered here showed progressive improvement in the quality of the fit, in the residuals and in the predictive skill. A T-only model had a number of shortcomings due to strong depth and latitudinal biases, but it may be the only option when no other properties are measured. For localized studies with restricted depth ranges (e.g., in the nearshore) a T-only model is probably adequate, but for ecosystem-wide studies where different water masses are expected to occur, the inclusion of proxy variables containing the spatial structure underlying these processes may help alleviate some of these problems (alternatively, separate T-only models could be fitted to stratified subsets of the data). A model that included salinity in addition to temperature corrected much of the latitudinal and depth biases. Finally, a model with oxygen, temperature, and salinity provided the most unbiased predictions, indicating that, at a minimum, these three variables are necessary to adequately describe the spatial and temporal processes driving nitrate variability throughout the CCS. Although it is possible to use density as a variable that combines temperature and salinity, modeling nitrate explicitly as a function of T and S had the added benefit of allowing us to visualize water masses in TS space, which was helpful in understanding the spatial distribution of nitrate. For studies where this is not necessary, modeling nitrate as a function of density would be more parsimonious from a modeling perspective. [55] Given the strong covariation of nitrate and temperature in the water column, it is not surprising that all the models explored had a good agreement between the response and the fitted values (D2  71.6–96.6%; Table 5), as has been widely reported in the literature. Similarly, the overall ability of the GAMs to predict nitrate observations

[58] Acknowledgments. This study was initially motivated by discussions with Frank Schwing and Roy Mendelssohn. The LTOP nutrient data were collected by Pat Wheeler (College of Oceanic and Atmospheric Sciences, Oregon State University) and are distributed through the U.S. GLOBEC Data System. John Largier (Bodega Marine Laboratory, University of California at Davis) and Richard Dugdale and Al Marchi (Romberg Tiburon Center, San Francisco State University) kindly provided access to the CoOP/WEST data set. William T. Peterson (NOAA/NMFS, Northwest Fisheries Science Center) and Jay Peterson (Cooperative Institute for Marine Resource Studies, Oregon State University) generously contributed the NH-25 data to this project. We thank Xuemei Qiu for programming assistance. Support came from a grant from the U.S. GLOBEC Northeast Pacific program and from the NOAA Fisheries and the Environment (FATE) program. This is U.S. GLOBEC contribution No. 729. The manuscript benefited from comments by Andrew W. Leising, Paul C. Fiedler, Richard C. Dugdale, and two anonymous reviewers.

References Bakun, A. (1973), Coastal upwelling indices, west coast of North America, 1946–71, U.S. Department of Commerce, NOAA Tech. Report, NMFSSSRF-671, Natl. Oceanic Atmos. Admin.,Washington, D. C.

16

PALACIOS ET AL.: MODELING THE T-N RELATIONSHIP Bjorkstedt, E. P., et al. (2011), State of the California Currrent, 2010– 2011: Regionally variable responses to a strong (but fleeting?) La Ni~ na, CalCOFI Rep., 52, 36–68. Bograd, S. J., D. A. Checkley, and W. S. Wooster (2003), CalCOFI: A half century of physical, chemical, and biological research in the California Current System, Deep Sea Res. Part II, 50(14-16), 2349–2353, doi:10.1016/S0967-0645(03)00122-X. Boyer, T. P., et al. (2009), World Ocean Database 2009, edited by S. Levitus, NOAA Atlas NESDIS 66, pp. 216, U.S. Gov. Print. Off., Washington D. C. Broitman, B. R., and B. P. Kinlan (2006), Spatial scales of benthic and pelagic producer biomass in a coastal upwelling ecosystem, Mar. Ecol. Prog. Ser., 327, 15–25, doi:10.3354/meps327015. Chavez, F. P., S. K. Service, and S. E. Buttrey (1996), Temperature-nitrate relationships in the central and eastern tropical Pacific, J. Geophys. Res., 101(C9), 20,553–20,563, doi:10.1029/96JC01943. Checkley, D. M., and J. A. Barth (2009), Patterns and processes in the California Current System, Progr. Oceanogr., 83, 49–64, doi:10.1016/ j.pocean.2009.07.028. Doney, S. C., I. Lima, J. K. Moore, K. Lindsay, M. J. Behrenfeld, T. K. Westberry, N. Mahowald, D. M. Glover, and T. Takahashi (2009), Skill metrics for confronting global upper ocean ecosystem-biogeochemistry models against field and remote sensing data, J. Mar. Syst., 76(1-2), 95– 112, doi:10.1016/j.jmarsys.2008.05.015. Dugdale, R. C., A. Morel, A. Bricaud, and F. P. Wilkerson (1989), Modeling new production in upwelling centers: A case study of modeling new production from remotely sensed temperature and color, J. Geophys. Res., 94(C12), 18,119–18,132, doi:10.1029/JC094iC12p18119. Dugdale, R. C., C. O. Davis, and F. P. Wilkerson (1997), Assessment of new production at the upwelling center at Point Conception, California, using nitrate estimated from remotely sensed sea surface temperature, J. Geophys. Res., 102(C4), 8573–8585, doi:10.1029/96JC02136. Friedland, K. D., C. Stock, K. F. Drinkwater, J. S. Link, R. T. Leaf, B. V. Shank, J. M. Rose, C. H. Pilskaln, and M. J. Fogarty (2012), Pathways between primary production and fisheries yields of large marine ecosystems, PLoS One, 7(1), e28945, doi:10.1371/journal.pone.0028945.t003. Fuentes, M., B. Xi, and W. S. Cleveland (2011), Trellis display for modeling data from designed experiments, Stat. Anal. Data Min., 4, 133–145, doi:10.1002/sam.10102. Garside, C., and J. C. Garside (1995), Euphotic-zone nutrient algorithms for the NABE and EqPac study sites, Deep Sea Res. Part II, 42(2-3), 335–347, doi:10.1016/0967-0645(95)00026-M. Goes, J. I., T. Saino, H. Oaku, and D. L. Jiang (1999), A method for estimating sea surface nitrate concentrations from remotely sensed SST and chlorophyll a—A case study for the North Pacific Ocean using OCTS/ ADEOS data, IEEE Trans. Geosci. Remote Sens., 37(3), 1633–1644, doi:10.1109/36.763279. Goes, J. I., T. Saino, H. Oaku, J. Ishizaka, C. S. Wong, and Y. Nojiri (2000), Basin scale estimates of sea surface nitrate and new production from remotely sensed sea surface temperature and chlorophyll, Geophys. Res. Lett., 27(9), 1263–1266, doi:10.1029/1999GL002353. Hastie, T. J., and R. J. Tibshirani (1990), Generalized Additive Models, Monographs on Statistics and Applied Probability 43, Chapman and Hall/CRC, Boca Raton, Fla. Henson, S. A., R. Sanders, J. T. Allen, I. S. Robinson, and L. Brown (2003), Seasonal constraints on the estimation of new production from space using temperature-nitrate relationships, Geophys. Res. Lett., 30(17), 1912, doi:10.1029/2003GL017982. Huyer, A., P. A. Wheeler, P. T. Strub, R. L. Smith, R. Letelier, and P. M. Kosro (2007), The Newport line off Oregon–Studies in the North East Pacific, Prog. Oceanogr., 75(2), 126–160, doi:10.1016/j.pocean.2007.08.003. Ihaka, R., and R. Gentleman (1996), R: A language for data analysis and graphics, J. Comput. Graphical Stat., 5, 299–314, doi:10.1080/ 10618600.1996.10474713. Kamykowski, D., and S. -J. Zentara (1986), Predicting plant nutrient concentrations from temperature and sigma-t in the upper kilometer of the world ocean, Deep Sea Res., 33(1), 89–105, doi:10.1016/0198-0149(86)90109-3. Kamykowski, D., S. -J. Zentara, J. M. Morrison, and A. C. Switzer (2002), Dynamic global patterns of nitrate, phosphate, silicate, and iron availability and phytoplankton community composition from remote sensing data, Global Biogeochem. Cycles, 16(4), 1077, doi:10.1029/2001GB001640. Lagerloef, G. (2012), Satellite mission monitors ocean surface salinity. Eos Trans. AGU, 93(25), 233–234. Louanchi, F., and R. G. Najjar (2000), A global monthly climatology of phosphate, nitrate, and silicate in the upper ocean: Spring-summer

export production and shallow remineralization, Global Biogeochem. Cycles, 14(3), 957–978, doi:10.1029/1999GB001215. Marra, G., and S. N. Wood (2011), Practical variable selection for generalized additive models, Comput. Stat. Data Anal., 55 (7), 2372–2387, doi:10.1016/j.csda.2011.02.004. McClatchie, et al. (2009), State of the California Currrent, spring 2008– 2009: Cold conditions drive regional differences in coastal production, CalCOFI Rep., 50, 43–68. Morin, P., M. V. M. Wafar, and P. Le Corre (1993), Estimation of nitrate flux in a tidal front from satellite-derived temperature data, J. Geophys. Res., 98(C3), 4689–4695, doi:10.1029/92JC02445. Palacios, D. M., S. J. Bograd, R. Mendelssohn, and F. B. Schwing (2004), Long-term and seasonal trends in stratification in the California Current, 1950–1993, J. Geophys. Res., 109, C10016, doi:10.1029/2004JC002380. Pe~ na, M. A., and S. J. Bograd (2007), Time series of the Northeast Pacific, Prog. Oceanogr., 75(2), 115–119, doi:10.1016/j.pocean.2007.08.008. Powell, T. M., C. V. W. Lewis, E. N. Curchitser, D. B. Haidvogel, A. J. Hermann, and E. L. Dobbins (2006), Results from a three-dimensional, nested biological-physical model of the California Current System and comparisons with statistics from satellite imagery, J. Geophys. Res., 111, C07018, doi:10.1029/2004JC002506. R Core Team (2012), R: A Language and Environment for Statistical Computing, R Found. for Stat. Comput., Vienna. [Available at http://www.Rproject.org/, accessed on 3 June 2013.]. Roy, C. (1991), Relations entre sels nutritifs et chlorophylle: Une approche exploratoire, in P^echeries Ouest Africaines: Cariabilite, Instabilite et Changement, edited by P. Cury and C. Roy, pp. 390–394, Editions de l’ORSTOM, Paris. Saba, V. S., et al. (2011), An evaluation of ocean color model estimates of marine primary productivity in coastal and pelagic regions across the globe, Biogeosciences, 8(2), 489–503, doi:10.5194/bg-8–489-2011. Sarangi, R. K. (2011), Remote-sensing-based estimation of surface nitrate and its variability in the southern peninsular Indian waters, Int. J. Oceanogr., Article ID 172731, doi:10.1155/2011/172731. Sarmiento, J. L., and N. Gruber (2009), Ocean Biogeochemical Dynamics, Princeton Univ. Press, Princeton, N. J. Sathyendranath, S., T. Platt, E. P. W. Horne, W. G. Harrison, O. Ulloa, R. Outerbridge, and N. Hoepffner (1991), Estimation of new production in the ocean by compound remote sensing, Nature, 353(6340), 129–133, doi:10.1038/353129a0. Schwing, F., M. O’Farrell, J. Steger, and K. Baltz, (1996), Coastal upwelling indices, west coast of North America, 1946–1995, U.S. Dep. of Commerce, NOAA Tech. Memo. NOAA-TM-NMFS-SWFSC-231, La Jolla, Calif. Sili o-Calzada, A., A. Bricaud, and B. Gentili (2008), Estimates of sea surface nitrate concentrations from sea surface temperature and chlorophyll concentration in upwelling areas: A case study for the Benguela system, Remote Sens. Environ., 112(6), 3173–3180, doi:10.1016/j.rse.2008.03.014. Somes, C. J., A. Schmittner, E. D. Galbraith, M. F. Lehmann, M. A. Altabet, J. P. Montoya, R. M. Letelier, A. C. Mix, A. Bourbonnais, and M. Eby (2010), Simulating the global distribution of nitrogen isotopes in the ocean, Global Biogeochem. Cycles, 24, GB4019, doi:10.1029/2009GB003767. Steinhoff, T., T. Friedrich, S. E. Hartman, A. Oschlies, D. W. R. Wallace, and A. Körtzinger (2010), Estimating mixed layer nitrate in the North Atlantic Ocean, Biogeosciences, 7(3), 95–807, doi:10.5194/bg-7–795-2010. Switzer, A. C., D. Kamykowski, and S. -J. Zentara (2003), Mapping nitrate in the global ocean using remotely sensed sea surface temperature, J. Geophys. Res., 108(C8), 3280, doi:10.1029/2000JC000444. Tabachnik, B. G., and L. S. Fidell (1989), Using Multivariate Statistics, Harper and Row, New York, N. Y. Tabachnik, B. G., and L. S. Fidell (2001), Using Multivariate Statistics, 4th ed., Allyn and Bacon, Boston, Mass. Traganza, E. D., V. M. Silva, D. M. Austin, W. L. Hanson, and S. H. Bonsink (1983), Nutrient mapping and recurrence of coastal upwelling centers by satellite remote sensing: Its implication to primary production and the sediment record, in Coastal Upwelling, Part A, edited by E. Suess and J. Thiede, pp. 61–83, Plenum, New York, N. Y. Wessel, P., and W. H. F. Smith (1996), A Global self-consistent, hierarchical, high-resolution shoreline Database, J. Geophys. Res., 101(B4), 8741–8743, doi:10.1029/96JB00104. Wood, S. N. (2006), Generalized Additive Models: An introduction with R, in Texts in Statistical Science, Chapman and Hall/CRC, Boca Raton, Fla. Zuur, A. F., E. N. Ieno, and C. S. Elphick (2009), A protocol for data exploration to avoid common statistical problems, Methods Ecol. Evol., 1(1), 3–14, doi:10.1111/j.2041-210X.2009.00001.x.

17